Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Fire Safety Journal 46 (2011) 451–461

Contents lists available at ScienceDirect

Fire Safety Journal


journal homepage: www.elsevier.com/locate/firesaf

Characterization of thermal properties and analysis of combustion behavior


of PMMA in a cone calorimeter
Jocelyn Luche a,n, Thomas Rogaume a, Franck Richard a, Eric Guillaume b
a
Institut Pprime, UPR 3346 CNRS, Département Fluides, Thermique, Combustion, ENSMA, BP 40109, 86961 Futuroscope, France
b
Laboratoire national de métrologie et d’essais (LNE), Test Direction, 78197 Trappes Cedex, France

a r t i c l e i n f o abstract

Article history: This paper deals with the thermal degradation of a black poly(methyl)methacrylate (PMMA) in a cone
Received 10 January 2011 calorimeter (CC) in air with a piloted ignition. The influence of several heat fluxes (11 kW m  2 and
Received in revised form 12 kW m  2, and ten values from 15 to 60 kW m  2 in steps of 5 kW m  2) on PMMA sample degradation
14 July 2011
and the decomposition chemistry has been studied. Thus, thermal properties have been deduced and
Accepted 18 July 2011
calculated from ignition time and mass loss rate (MLR) curves. During our experiments, among
Available online 9 August 2011
compounds quantified simultaneously by a Fourier transformed infrared (FTIR) or gas analyzer, five
Keywords: main species (CO2, CO, H2O, NO and O2) have been encountered, regardless of the external heat flux
PMMA considered. The main product concentrations allow calculation of the corresponding emission yields.
Cone calorimeter
Thus, mass balances of C and H atoms contained in these exhaust gases were able to be compared with
FTIR
those included in the initial PMMA sample. Using the standard oxygen consumption method, heat
Heat release rate
Mass loss rate release rate (HRR), total heat release (THR) and effective heat of combustion (EHC) have been calculated
Emission yields for each irradiance level. Therefore, these different results (thermal properties, emission yields, HRR,
Effective heat of combustion THR and EHC) are in quite good accordance (same order of magnitude) with those found in previous
Thermal properties studies.
ISO 5660 standard & 2011 Elsevier Ltd. All rights reserved.

1. Introduction fire, as under the right conditions, they readily ignite and burn
vigorously.
Plastics and polymers, such as poly(methyl)methacrylate As a reference fuel material used in a cone calorimeter (CC),
(PMMA), are used as replacements for conventional materials solid acrylic PMMA polymer has been widely used – with or
(e.g. wood, metals, glass, etc.) in numerous domains such as without filler – during previous studies for assessing polymer
electrical appliances, light diffusers, optical fibers, furniture, flammability and characterizing the mass loss rate (MLR) during
transport, hygiene and health, etc. Indeed, these widespread kinds combustion processes [1–14]. Only some of these previous
of synthetic compounds, with good technical, mechanical, chemi- studies and other specific studies propose the chemical analysis
cal, optical performances, etc., are easily manufactured for a and quantification of the effluents emitted during PMMA thermal
moderate cost and can be used in a wide range of applications. degradation, and so include the calculation of the heat release
Thus, compounds made of PMMA can be used in numerous rate (HRR), which is the most important parameter in fire
applications as signs and signboards (illuminated panels, 3D characterization studies [1–5,15–22]. Finally, a few studies pro-
lettering, indicator panels, etc.), POS (point-of-sale) advertising pose a chemical kinetic mechanism for predicting the PMMA
(display stands, testers, notice-boards, etc.), interior design (shop- degradation in a CC [23,24]. Generally, only the assumption of a
fitting, furniture, projection screens, glazing, etc.), transport single-step degradation corresponding to depolymerization and
(deflectors, sun visors, registration plates, ship portholes and formation of MMA is carried out.
windows, etc.) or industrial (machine guards, dials, precision The present paper deals with a complete and detailed study of
parts, etc.) devices. In spite of these various uses linked to thermal degradation of a black non-charring PMMA in a CC,
numerous advantages, plastics (highly combustible materials) including the characterization of thermal properties and the
constitute a grave danger (human injuries and deaths) in case of quantification of exhaust gas concentrations of this typical poly-
mer. Thus, the main purpose of this work is to characterize the
influence of a widespread range of irradiance levels (external heat
flux) on the MLR of a PMMA sample and on the product amounts
n
Corresponding author. released during this combustion process. Next, the HRR was
E-mail address: jocelyn.luche@lcd.ensma.fr (J. Luche). determined based on the O2-consumption principle in a CC.

0379-7112/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.firesaf.2011.07.005
452 J. Luche et al. / Fire Safety Journal 46 (2011) 451–461

Nomenclature HRR averaged heat release rate, kW m  2


k thermal conductivity, kW m  1 K  1
Greek notations
m mass, g
MAir molar mass of air, g mol  1
e emissivity m_ Air mass flow rate of the incoming air, g s  1
f oxygen depletion factor Mi molar mass of species i, g mol  1
krc thermal effusivity, kJ2 m  4 K  2 s  1 m_i mass flow rate of species i in the exhaust duct of CC,
r density, kg m  3 g s1
s Stefan–Boltzmann constant 5.6704  10  8 W m  2 K  4 m_e mass flow rate in the exhaust duct of CC, g s  1
MLR mass loss rate, g s  1
Notations NDIR non-dispersive infrared, –
P pressure, Pa
A sample area exposed to cone calorimeter heat flux, PMMA poly(methyl)methacrylate, –
cm2 q_ 00ext external heat flux, kW m  2
c specific heat, kJ kg  1 K  1 R universal constant of perfect gas
CC cone calorimeter, – 8.314472 J mol  1 K  1
CHF critical heat flux, kW m  2 SMLR s mass loss rate, g m  2 s  1
DHg latent heat of gasification, kJ g  1 SMLR averaged specific mass loss rate, g m  2 s  1
E net heat release per unit mass of oxygen consumed, t time, min or s
MJ kg  1 of O2 T temperature, K
ECO net heat release per unit mass of oxygen for CO, tig ignition time, min or s
MJ kg  1 of O2 Tig ignition temperature, K
EHC effective heat of combustion, kJ g  1 T0 initial (ambient) temperature, K
EHC averaged effective heat of combustion, kJ g  1 THR total heat release, MJ m  2
FHF flame heat flux, kW m  2 V_ e volumetric flow rate in the exhaust duct, L s  1
FHFnet net flame heat flux, kW m  2 Vm molar volume, L mol  1
FID flame ionization detector, – Xi0 mole fraction of species i in the incoming air, –
FTIR Fourier transformed infrared, – Xi mole fraction of species i in the exhaust gas measured
hc convective heat transfer coefficient, W m  2 K  1 by analyzer, –
HRR heat release rate, kW m  2 Yi emission yield of species i, gi/gsample

Finally, the MLR and HRR parameters have been correlated with Elementary analysis results show that no inert load, flame
main gaseous compound evolutions and their emission yields to retardants or fillers were used during the manufacturing of the
propose a description of physical and chemical phenomena PMMA sample; neither chlorine- nor sulfur-based additives were
occurring during PMMA thermal degradation. found. Indeed, 100 wt% of the total sample mass was composed of
C, H, O, N and S atoms. This result is in good accordance with
various compositions of such PMMA reported in [25]. Based on
this elementary analysis composition, the raw chemical formula
2. Experiments of the virgin PMMA was determined to be (C4.9H7.8O2.0)n (with
n¼PMMA polymerization degree).
2.1. Materials Its molecular weights and molecular weight distributions were
obtained from an Agilent Technologies Size Exclusion Chromato-
The material used in this study is a black non-charring PMMA, graphy (SEC) calibrated with a polystyrene standard. Samples of
commonly known as Altuglas, supplied by the company VACOUR 25 and 70 mg (1 and 2 in Table 2) were dissolved in chloroform
and synthesized via radical polymerization. The PMMA elemen- and then injected into the SEC. Table 2 shows the molecular
tary analysis was conducted by a combination of catharometry weights (Mn and Mw), molecular weight distributions (Ip) and
and non-dispersive infrared (NDIR) detection. Table 1 presents the polymerization degree (n) of each sample analyzed.
averaged composition along with the analysis methods that were Molecular weight distribution (Ip), with a value around 3.7,
repeated three times and provided a dispersion of 70.3 wt%. shows that PMMA is a polymolecular compound obtained by
radical polymerization (IpZ2) according to the definition given
Table 1 by some authors in the literature [26,27]. From these results, the
Elemental analysis of black poly(methyl
methacrylate).
polymerization degree (n) of the PMMA sample can be deter-
mined as equal to value ranging from 1500 to 1800.
Elements Composition (wt%)

Carbon (C) 59.1


Hydrogen (H) 7.9 Table 2
Oxygen (O) 31.9 Characterization of black poly(methyl)methacrylate mass (SEC analysis).
Nitrogen (N) o0.3
Sulfur (S) o0.2 Elements Mn Mw Ip¼ Mw/Mn n
Chlorine (Cl) 0.1
Water (H2O) 0.6 1 178,670 639,340 3.6 1785
Total o100.1 2 150,440 578,480 3.8 1503
J. Luche et al. / Fire Safety Journal 46 (2011) 451–461 453

2.2. Cone calorimeter HORIBA apparatus during our experiments. Moreover, this
FTIR spectrometer allows the identification and quantification
The PMMA reaction-to-fire characterization was carried out in of other combustion products such as NO2, NH3, HCN, N2O,
a CC (Standard ISO 5 660 [28]) made by Fire Testing Technology CH4, C2H2, C2H4, C2H6, C3H6, C3H8 and H2O (not quantified by
Limited, under fully ventilated conditions. The sample dimensions the HORIBA gas analyzer).
were 99 71 mm long, 9971 mm wide and 14 71 mm thickness,
with a mass of 169.6 72.9 g. Thus, mass density (r) calculated
Furthermore, before all the experiments, the two devices used
from these data is found to be equal to 1213.3719.1 kg/m3. Fig. 1
for gas analysis were calibrated with well-known concentration
presents a schematic view of the CC with a solid PMMA matrix in
standards to quantify the five gaseous combustion products for
the sample holder exposed to an irradiance level from an electric
the HORIBA apparatus and the 15 products for the FTIR spectro-
heater. Several heat flux levels were used: 11 and 12 kW m-2, and
meter measured simultaneously during the combustion process.
ten heat flux levels from 15 to 60 kW m-2 in steps of 5 kW m-2.
The entire transport line (from the sampling point to the
Tests were carried out with a piloted ignition in air, and were
quantification apparatus) was heated to 170 1C and was not dried
repeated at least five times for each condition. Moreover, the
before passing through the FTIR measurement gas cell. This
ignition spark was positioned above the sample up to ignition
allowed the transport of combustion products while avoiding
(and removed thereafter), so all the tests were performed in a
water vapor condensation and the trapping of water-soluble
flaming condition. According to the standard ISO 17554 [29], the
compounds. It also enabled quantification of H2O (vapor).The
experiments were stopped manually if no ignition occurred after
sampling line was connected to a filtration box where a cellulosic
30 or 32 min after ignition or when the mass loss became zero.
filter and a stainless steel filter were used to retain heavy
During these experiments, the CC fan flow rate was taken to be
products and soot particles. After filtration, gases were trans-
equal to 0.024 70.002 m3 s-1. At the end of all experiments and,
ported up to the gas analyzers (FTIR and HORIBA). The FTIR
regardless of the external heat flux chosen, all the PMMA sample
analysis technique used (including sampling and filtering device)
was entirely degraded, and no solid or liquid residue was found
was validated during the SAFIR project [30], which constituted
inside the sample holder.
the basis for toxicity analysis carried out following the guidelines
of the standard ISO 19702 [31–35].
2.3. Gas analysis

Exhaust gases produced by the thermal degradation of the


PMMA sample in the CC were sampled from the exhaust duct 3. Thermal property results
using two kinds of on-line gas analyzers:
3.1. Parameters of inflammation and combustibility
 A gas analyzer HORIBA PG 250 equipped with three units to
simultaneously quantify gaseous products (NO, O2, CO2, CO For the PMMA studied in the CC under the conditions outlined
and SO2). The NOx analysis unit uses a cross-modulation in the previous part, sample ignition can be achieved when an
ordinary pressure chemiluminescence method, the SO2, CO, external heat flux q_ 00ext (with a relative uncertainty of 75%), higher
and CO2 unit uses an NDIR absorption method and the O2 unit than a critical heat flux (CHF), is applied during a time interval tig .
uses an electrochemical Zirconia method using a galvanized These different parameters can be defined and calculated by using
cell. The sampling flow rate was 0.4 NL min  1. Eqs. 1 and 2, which are described and fully detailed in Refs.
 A Fourier transformed infrared (FTIR) spectrometer (Thermo- [1–4,14]:
Nicolet 6700 equipped with a MCT-A detector and a measure-  
ment gas cell of volume¼0.2 L and optical path-length¼2 m). 2 Tig T0 2
tig ¼ krc 00 ð1Þ
The sampling flow rate was equal to 2 NL min  1. The FTIR 3 q_ ext
analyzer was also used to quantify the NO, CO2, CO and SO2
species, which allows the control and comparison of these 1h i
q_ 00ext ¼ hc ðTig T0 Þ þ esTig4  CHF ð2Þ
concentration values with those values measured using the e

Cone calorimeter fan

Spark ignition

Temperature and differential


pressure measurements

Cone calorimeter hood Soot filter

GAS ANALYZERS:
-FTIR
-HORIBA PG250

Heated sampling line (170°C)

Cone calorimeter (Electrical heater)

Weighting Sample holder


device

Fig. 1. Schematic layout of the coupling of cone calorimeter and gas analyzers.
454 J. Luche et al. / Fire Safety Journal 46 (2011) 451–461

" # " #
1 e 00
hc ðTig T0 Þ þ esTig4 difference between theoretical and experimental values of CHF
pffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi _
Uq ext  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð3Þ has suggested the influence of the surface reradiation of the
tig ð2=3ÞkrcðTig T0 Þ ð2=3ÞkrcðTig T0 Þ
PMMA sample, which causes a higher experimental CHF than
1=2
The slope of a plot obtained from the tig ¼ f ðq_ 00ext Þ curve the value extrapolated from Fig. 2. Moreover, different methods
(Eq. (3)) allows computation of the different thermal properties described in previous studies [37–39] have shown that the
such as the theoretical critical heat flux (CHF), the specific heat theoretical CHF value can be used to calculate a more realistic
(c), the thermal conductivity (k) and the theoretical ignition CHF. Indeed, for example, some intercept values (CHF) are defined
temperature (Tig) by using the following equations: by the authors as being only 0.64 [37] or 0.76 [38] fraction of
" # " # the CHF.
4
e hc ðTig T0 Þ þ esTig
Slope ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; yintercept ¼  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð2=3ÞkrcðTig T0 Þ ð2=3ÞkrcðTig T0 Þ
3.2. Transient burning rate
ð4Þ
  Fig. 3 shows the experimental transient evolution of the MLR
yintercept
CHF ¼  ð5Þ at the different external irradiance levels or external heat flux
Slope
used during the PMMA thermal degradation experiments. Time
Fig. 2 shows plots of the curve evolution of the square root of t¼0 marks the beginning of the exposure to the desired irra-
the inverse of ignition time as a linear function of external diance level rather than the moment of ignition.
irradiance (Eq. (3)), with a straight line used to fit the data Regardless of the heat flux chosen, three or four – more or less
(slope¼0.00509 and yintercept ¼  0.02434) up to the minimum significant – decomposition time ranges (corresponding to three
experimental heat flux corresponding to critical ignition or four peaks or slope variations, which reach a maximum MLR
(11 kW m  2) represented by a vertical dashed line. Moreover, in value as summarized in Table 3) can be observed in Fig. 3 with a
Fig. 2, the fit line (the curve slope) reaches an intersection point major one that has the most significant peak intensity (corre-
with the x-axis of approximately 4.8 kW m  2. This value repre- sponding to the 3rd stage in Table 3). Moreover, an increase of
sents the theoretical CHF for PMMA ignition computed from external heat flux moves the different stages towards the first one
Eqs. (4) and (5). This CHF theoretical result is in good accordance while increasing curve intensity. Indeed, when the external heat
with previous results found in the literature where the CHF flux increases from 11 to 60 kW m  2, all the PMMA sample mass
(obtained by the same way) is equal to 4 kW m  2 [11] or is lost on a shorter time (e.g. around 10 min at 60 kW m  2 against
5 kW m  2 [1–5]. The effective CHF determined experimentally 45 min at 11 kW m  2), with a higher maximal intensity of the
in this study (11 kW m  2) is in good accordance with results MLR peak (e.g. around 0.5 g s  1 at 60 kW m  2 compared to
found in the literature (9–13 kW m  2 [5,11,14,36]). 0.15 g s  1 at 11 kW m  2). As shown, the MLR shapes and their
The difference between the experimental and extrapolated CHF intensities depend strongly on and change with the irradiance
is due to the non-linearity of the ignition time leading to the level value. After a latency time, the first stage starts with the
non-linearity of the square root of the inverse of ignition time ignition of the PMMA sample and the very quick rise of MLR after
(Fig. 2) for the lowest external heat flux, especially near the ignition. Next, the PMMA decomposition continues during the
CHF [14] (where the solid material is thermally thin compared 2nd and 3rd time ranges, which correspond to the MLR evolution
to the time of thermal transfer). Another explanation [5] of this following a plateau (2nd column of Table 3) and reach a
significant peak (3rd column of Table 3), respectively. Finally,
the last decomposition time range (4th column of Table 3)
corresponds to an MLR decrease and the consumption of the
remaining PMMA amount in the sample holder. At the end of the
different experiments, no liquid or solid residue was found in the
sample holder.
The specific mass loss rate (SMLR) is determined as the ratio
between the MLR and the sample surface exposed to the CC
external irradiance level (i.e. A¼88.4 cm2, as explained in the
paragraph 6.7 of the ISO 5660 standard [28], where the opening
for the specimen face exposed to heat flux coming from the CC
heater is equal to 9.470.5 cm  9.470.5 cm). Moreover, the
SMLR is defined (Eq. (6)) as the sum of the heat flux from the
flame (FHF) and the external heat flux (q_ 00ext ), less the radiative
heat flux loss (esTig4 ), divided by the latent heat of gasification
(DHg ). According to a hypothesis carried out in previous studies
[1–4], black PMMA approximates a vaporizing solid, and the
flame heat flux (FHF) is constant. The FHF [1–4] or total
FHF [9] is defined as the sum of the radiative (FHFr) and
convective (FHFc) components of the FHF (FHF ¼ e FHFr þFHFc ,
where e is emissivity). The net FHF [1–4,9] is defined as the
material’s contribution to burning in the CC and is the difference
between (total) FHF and radiative heat flux loss from the sample
surface (FHFnet ¼ FHFesTig4 ).
  4
1 FHFesTig
SMLR ¼ q_ 00ext þ ð6Þ
DHg DHg

Fig. 2. Square root of the inverse of ignition time (s  0.5) as a function of heat flux Experimental averaged SMLRs are plotted in Fig. 4 as functions
(kW m  2). of external incident heat flux. The SMLR (or MLR) is accelerated
J. Luche et al. / Fire Safety Journal 46 (2011) 451–461 455

Table 3
Time range of the four stages of PMMA thermal degradation.

Time range (min)

Heat flux (kW m  2) 1st stage 2nd stage 3rd stage 4th stage

11 23  25 25  35 35  45 445
12 13  15 15  24 24  34 434
15 57 7  15 15  25 425
20 23 3  12 12  18 418
25 12 29 9  15 415
30 0.5  1.5 1.5  8 8  14 414
35 0.5  1 18 8  11.5 411.5
40 0.5  1 1  7.5 7.5  10.5 410.5
45 0.2  0.4 0.4  7 7  10 410
50 0.2  0.4 0.4  6.5 6.5  9 49
55 0.2  0.4 0.4  5 5  8.5 48.5
60 0.2  0.4 0.4  5 58 48

Fig. 4. Averaged specific mass loss rate SMLR (g m  2 s  1) vs. heat flux (kW m  2).

The slope ( ¼0.4268) and yintercept ( ¼6.3606) of the best fit line


to flaming black PMMA data obtained from the SMLR ¼ f ðq_ 00ext Þ
curve (Eq. (6)) allows the computation of other thermal properties
such as gasification heat (DHg ¼2.34 kJ g  1), flame heat flux
(FHF¼17.67 kW m  2) and net flame heat flux (FHFnet ¼
14.90 kW m  2) by using Eq. (7), which is deduced from Eq. (6):

1 FHFesTig4
Slope ¼ ; yintercept ¼ ð7Þ
DHg DHg

The black PMMA gasification heat (DHg ¼2.34 kJ g  1) found in


this study is in good accordance with previous results from the
literature, as presented in Table 4.
As seen for thermal properties and CHF, theoretical results
calculated or experimental ones measured during this black
Fig. 3. Mass loss rate (g s  1) as a function of experimental time (min).
PMMA thermal degradation study in a CC are of the same
magnitude order as those found previously in literature. Slight
when the irradiance level is increased up to a high value. Indeed, differences can be observed between the different results; this is
the SMLR increases from 11 to 32 g m  2 s  1 when incident heat probably due to the wide disparity between PMMA samples used,
flux increases from 11 to 60 kW m  2. indicating a real difference due to elementary composition and
456 J. Luche et al. / Fire Safety Journal 46 (2011) 451–461

density as well as experimental measurement uncertainties. The Me ¼ 18 þ 4ð1XH2 O ÞðXO2 þ 4XcO2 þ 2:5Þ ð11Þ
difference between results can also be explained by the difference
due to the cone designs used or due to result interpretations. In where A¼88.4 cm2 (sample area exposed to CC heat flux),
1
the latter case, in the literature for theoretical CHF determination, E¼13.1 MJ kg  1 of O2; ECO ¼17.6 MJ kg  1 of O2; MO2 ¼ 32 g mol ,
only a part of the results (up to 40 kW m  2 in [1–4] or between MAir ¼29 g mol  1, m_ Air is the mass flow rate of the incoming air, m
_e
20 and 50 kW m  2 in [5]) are fitted by a straight line, which has a is the mass flow rate in the exhaust duct of the CC, Xi is the mole
more or less significant influence on the CHF determination, and fraction of species i in the exhaust gas measured by the analyzer, Xi0
so influences the computation of other parameters (c, k, DHg). is the mole fraction of species i in the incoming air and f is the
oxygen factor.
Fig. 5 shows the transient evolution of the SMLR and the HRR
3.3. Heat release rate (HRR) at four irradiance levels (15, 30, 45 and 60 kW m  2). As seen
previously for MLR, HRR and SMLR in this paper, transient
The HRR is the most significant parameter for the fire hazard evolutions of these parameters depend strongly on the irradiance
material evaluation [40,41] since it controls the rate of growth in level, and both parameters have the same curve shapes compared
fire, including heat and ultimately the amount of smoke and toxic to the MLR one for a given external heat flux. As for the MLR
gas generated. This parameter is measured by using the oxygen transient evolution in Fig. 3, the four PMMA thermal decomposi-
consumption calorimetry technique [42] based on the ISO 5660 tion stages summarized in Table 3 are also found from HRR and
standard [28]. The O2, CO2, CO, H2O and N2 species account for SMLR curves in Fig. 5, regardless of the incident heat flux applied
approximately 99% of the exhaust gases in the majority of full- to the sample.
scale fire tests [43,44]. Thus, these gaseous compounds, identified
as the major species in our experimental results (Fig. 7 and
Table 5), will be considered as the only species in the exhaust
gas flow. As H2O concentrations were measured during our
experiments, the HRR value can be computed by the following
equations (described in detail in Refs. [43–49]):
 
1 1f XCO MO2
HRR ¼ EfðECO EÞ _ ð1XH0 O XCO
m 0
ÞXO0 2 ð8Þ
A 2 XO2 MAir Air 2 2

XO0 2 ð1XCO XCO2 ÞXO2 ð1XCO


0
Þ
f¼ 2
ð9Þ
XO0 2 ð1XO2 XCO XCO2 Þ

_ Air
m _e
ð1XH2 O Þð1XO2 XCO XCO2 Þ m
¼ ð10Þ
MAir ð1XH0 2 O Þð1XO0 2 XCO
0 Þ
2
Me

Table 4
Comparison of PMMA gasification heat of this
study with values from literature.

DHg (kJ g  1) References

2.34 This study


1.34–1.48 [12]
1.6 [5,6,10,11–13]
1.96 [8]
2.2–2.7 [6]
2.6 [9]
2.77 [1–4]
Fig. 5. HRR and SMLR curves as functions of time at different external heat flux.

Table 5
HRR, SMLR , EHC and THR obtained during this study and from literature.

Results from this work Results from literature

2 2
Heat flux (kW m ) HRR (kW m 2
) SMLR (g m 2
s 1
) EHC (kJ g 1
) THR (MJ m ) HRR (kW m)  2 EHC (kJ g  1)

11 274.3 7 116.1 11.2 74.5 24.0 72.3 445.6


15 317.4 7 111.1 12.8 74.4 25.0 73.0 450.1 21.8 [23]
20 349.4 7 138.5 15.0 75.3 23.7 76.7 452.9
25 412.0 7 166.0 17.4 76.3 24.0 76.2 450.4 401.56 [1–4] 24.0 [1–4]
30 463.8 7 222.0 18.6 77.9 24.8 74.0 466.4
35 499.9 7 238.7 20.8 79.0 24.7 79.1 451.1 23.2 [15,16]
40 577.7 7 267.4 23.2 79.9 24.9 72.3 466.4
45 633.4 7 303.4 25.9 710.5 24.0 74.1 471.1
50 691.4 7 346.9 27.7 711.4 24.6 74.4 492.6 642.21 [1–4] 24.1 [1–4]
55 735.3 7 315.4 30.5 712.2 25.3 77.7 459.4
60 780.8 7 365.7 31.6 713.0 24.3 72.8 468.5
Mean 521.4 7 174.8 21.3 77.0 24.5 70.5 461.3 7 13.6 24–25
[5,14,17–21,45]
J. Luche et al. / Fire Safety Journal 46 (2011) 451–461 457

The effective heat of combustion (EHC) value is a time- and heat flux equal to 15, 30, 45 and 60 kW m  2 (Fig. 7a–d, respec-
irradiance-level-dependent parameter that corresponds to the tively). As can be seen in Fig. 7, an influence of the external heat
heat released from the volatile portion during solid material flux value can be noticed between the different curve sets for each
combustion. From the HRR and SMLR, the EHC can be computed exhaust gas. Indeed, when the irradiance level increases from 11
using the following equation: to 60 kW m-2, CO, CO2, H2O, NO, NO2 and HCN concentration
HRR levels increase, and the O2 one decreases.
EHC ¼ ð12Þ During these experiments, concentrations of CH4, C2H2, C2H4,
SMLR
C2H6, C3H6 and C3H8 have been measured with maximum values
Averaged results for SMLR, HRR, EHC and total heat release (THR)
that never exceed 10 ppm. No significant amounts ( o1 ppm) of
are presented in Table 5 for the set of irradiance levels studied (from
SO2, N2O or NH3 have been quantified (i.e. the maximum value
11 to 60 kW m  2). Globally, HRR increases from 275 to 780 kW m  2,
measured for each concentration is near the detection limit, but
SMLR increases from 11 to 32 g m  2 s  1, EHC values are around to
lower than the quantification limit of the gas analysis apparatus
24.570.5 kJ g  1 and THR values are around 460714 MJ m  2 with
for these gaseous compounds).
increasing external heat flux (from 11 to 60 kW m  2, respectively).
Moreover, curve evolutions of main exhaust gas production
These averaged HRR and/or EHC results are compared to some black
(and thus O2 consumption) are linked to the MLR curve evolution.
PMMA thermal degradation data found in the literature. As seen in
Indeed, when the PMMA sample is degraded (i.e. MLR increases),
Table 5, results found during this black PMMA thermal degradation
we can observe an oxygen consumption (O2 concentration
study in the CC are in quite good accordance (same order of
decreases) and simultaneously an exhaust gas production
magnitude) with those given by previous studies.
(concentrations of volatile gases, such as CO, CO2, H2O, NO, NO2
and HCN, increase). Conversely, when the PMMA thermal degrada-
4. Exhaust gas concentrations and emission yields tion becomes less important (i.e. MLR decreases), O2 amount rises
(i.e. reaches the initial ambient value of O2 concentration) and
In Fig. 6, CO2 concentrations obtained simultaneously from the other major product concentrations decrease. Thus, the curve
gas analyzer (HORIBA) and the FTIR spectrometer are plotted as shapes of these different products (and the oxygen one by the
functions of time for external heat fluxes equal to 15, 30, 45 and mirror effect) correspond perfectly to the MLR ones. So, the
60 kW m  2. As can be seen in Fig. 6, regardless of the heat flux physical and the chemical events arising during the PMMA thermal
considered, good accordance is observed between the CO2 degradation process can be identified by correlating the MLR and
concentrations measured with the two types of gas analysis gaseous compound curves with Table 3 given in Section 3.
devices. The same observations can be carried out for the CO and The product emission yields quantified during the different
NO concentration comparison between the HORIBA gas analyzer experiments are useful tools to extrapolate exhaust gaseous
and the FTIR spectrometer. Therefore, regardless of the CO or CO2 species production obtained from bench-scale (CC) to full-scale
concentrations chosen (measured with the HORIBA or the FTIR scenarios [42]. Table 6 gives the emission yields of the main
apparatus), this choice has a minor impact (o2%) on the different products released during the PMMA decomposition process. The
calculation results, i.e. on HRR, EHC, THR or emission yield results. gaseous species emission yields (Yi), calculated as the ratio
between the exhaust product i mass flow rate (m _ i ) and the PMMA
External heat flux influence on PMMA sample thermal degra-
dation has been studied with concentration evolution of major MLR, are expressed by the following equation:
products. Evolution curves of CO, CO2, O2, H2O, NO, NO2 and HCN m_i
concentrations are plotted as functions of time for an external Yi ¼ ð13Þ
MLR
The mass flow rate of species i released during the combustion
process (Eq. (14)) is represented by the product of the species
mole fraction (Xi), the volumetric flow rate in the exhaust duct
(V_ e ), the molar mass of species i (Mi) divided by the molar volume
(Vm) given by Eq. (15), where R is the universal constant of a
perfect gas, and T and P are the temperature and the pressure of
the gas mixture in the CC exhaust duct, respectively;

Xi V_ e Mi
_i¼
m ð14Þ
Vm

RT
Vm ¼ ð15Þ
P
During the burning process, only three species (CO, CO2 and
H2O) present emission yields of high consistency. Emission yields
of these species are found to be quite constant (taking into
account uncertainties on measurements and calculations), regard-
less of the initial irradiance level used during the experiments.
Indeed, in Table 6, CO and H2O emission yields decrease slightly
from 0.008 to 0.006 gCO/gsample for CO and from 0.7 to 0:6 gH2 O =g
for H2O when the irradiance level increases from 11 to
60 kW m  2, respectively. Conversely, CO2 emission yields
increase slightly from 2.0 to 2:2 gCO2 =g when the external heat
flux increases from 11 to 60 kW m  2, respectively. Moreover, for
the irradiance level range studied (from 11 to 60 kW m  2), mean
Fig. 6. Comparison of CO2 concentrations (ppm) from gas analyzer and FTIR as values are found to be equal to 0.007 gCO/gsample, 2:148 gCO2 =g
functions of time (min) at 15, 30, 45 and 60 kW m  2. and 0:658 gH2 O =g . Other chemical compounds (e.g. NOx, HCN or
458 J. Luche et al. / Fire Safety Journal 46 (2011) 451–461

Fig. 7. Major product concentrations (ppm) and mass loss rate (g s  1) as functions of time (min) at 15 (a), 30 (b), 45 (c) and 60 kW m  2 (d).

Table 6
Emission yields of main exhaust products released during PMMA combustion.

External heat Chemical compound yields (gspecies/gsample) CO/CO2 (%)


flux (kW m  2)
CO CO2 H2O

11 0.008 70.001 2.034 70.407 0.720 7 0.144 0.362


12 0.008 70.002 1.977 7 0.395 0.758 7 0.152 0.383
15 0.008 70.002 2.126 7 0.425 0.748 7 0.150 0.380
20 0.008 70.002 2.169 7 0.434 0.694 7 0.139 0.361
25 0.006 70.001 2.069 70.414 0.655 7 0.131 0.309
30 0.006 70.001 2.160 70.432 0.633 7 0.127 0.293
35 0.006 70.001 2.182 7 0.436 0.640 7 0.128 0.275
40 0.006 70.001 2.206 70.441 0.582 7 0.116 0.266
45 0.006 70.001 2.216 7 0.443 0.649 7 0.130 0.262
50 0.006 70.001 2.217 7 0.443 0.611 7 0.122 0.250
55 0.006 70.001 2.201 70.440 0.614 7 0.123 0.262
60 0.006 70.001 2.221 7 0.444 0.587 7 0.117 0.257
Mean 0.007 70.001 2.148 7 0.430 0.658 7 0.132 0.305

light hydrocarbons), which have been measured during these 5. Discussion


experiments, have low emission yield values (i.e. o1 m g/gsample),
regardless of the irradiance level considered. The thermal degradation study of virgin PMMA has allowed the
During this study in a CC, the averaged emission yield results computation of several thermal properties (Section 3), of this
found for CO2 and CO species (i.e. 2:148 gCO2 =gPMMA and polymer, which are in good accordance with previous studies:
0.007 gCO/gPMMA) are comparable with those found in previous theoretical CHF 5 kW m  2, experimental CHF¼ 11 kW m  2,
studies [1–4] (i.e.  2:54 gCO2 =gPMMA and  0.008 gCO/gPMMA), DHg ¼2.34 kJ g  1, averaged heat release rate HRR  275–780
[14] (i.e.  2:12 gCO2 =gPMMA and  0.010 gCO/gPMMA), [15,16] kW m  2 for external heat fluxes ranging from 11 to 60 kW m  2,
(i.e.  3:23 gCO2 =gPMMA and  0.016 gCO/gPMMA) or [40–42] respectively, and EHC 24.5 kJ g  1, regardless of the external heat
(i.e.  2:05 gCO2 =gPMMA and  0.010 gCO/gPMMA). flux chosen.
J. Luche et al. / Fire Safety Journal 46 (2011) 451–461 459

The chemical analysis (Section 4) of the exhaust gases place and the associated physical and chemical phenomena can be
by FTIR, NDIR and flame ionization detection (FID) has shown identified as follows:
that only four products (CO, CO2, H2O and NO) and O2 were
measured as main species. Indeed, low concentrations were – A first one (1st column in Table 3) corresponding to the start of
measured for the other compounds, such as NO2, light hydro- thermal degradation with the beginning of ignition. During
carbons and HCN. Emission yields of all species quantified during this first time range, a significant bubbling phenomenon at the
this work have been calculated and only CO, CO2, H2O, NO and O2 PMMA sample surface occurred, and all concentrations (except
have values of high consistency (i.e. higher than 1 mg/gsample). for the O2 one, which decreased) and MLR increased strongly.
This was confirmed by the atomic balances for C and H presented – A second one (2nd column in Table 3) corresponding to a thick
in Table 7, which compares the C and H atom masses contained in PMMA sample thermal degradation. During this second stage,
the initial PMMA and those contained in the main gaseous a swelling phenomenon of the PMMA sample occurred and all
products (i.e. CO, CO2 and H2O). Indeed, C atom masses from exhaust gas amounts and MLR continued increasing (except
CO and CO2, and H atom masses from H2O are, in the majority for the O2 concentration, which decreased), but more slowly
of cases (taking into account measurement uncertainties), than previously.
equal to the initial C and H atom masses of the PMMA sample, – A third one (3rd column in Table 3) corresponding to a thin
respectively. PMMA sample thermal degradation. During this stage, the
The results in Table 7 show that all initial C and H amounts swelling phenomenon ended, and PMMA matrix surface crack-
(taking into account measurement uncertainties) are converted ing occurred. Thus, O2 consumption and effluent production,
into H2O, CO and CO2 species, and no other compound (or in very as well as the MLR value, reach an optimum value, which
small amounts, i.e. o1 mg/gsample, as measured for light hydro- corresponds to the main peak (maximum for exhaust gases
carbons or HCN) containing high quantities of C and/or H atoms and minimum for oxygen) observed in Fig. 7.
(e.g. heavy hydrocarbons, polycyclic aromatic hydrocarbons, soot, – A fourth one (4th column in Table 3) corresponding to the
etc.) seems to be produced during thermal degradation of PMMA. oxidation of the remaining amount of the PMMA sample up to
Thus, the same amount of exhaust gas (i.e. the same value of the end of the degradation process (end of the ignition). During
exhaust gas emission yield) is produced during the thermal this time range, small CO, CO2 and other gas production
degradation process, regardless of the external heat flux applied (exhaust product concentrations and MLR values decrease)
to the PMMA sample (no residue found in the sample holder at and small O2 consumption (oxygen concentration increases up
the end of the different experiments). However, the MLR or HRR to the initial value) were observed. At the end of this step, no
correlated with species concentration evolutions have shown that residue was found in the sample holder.
external heat flux has an influence on the intensities of sample
degradation (transient MLR curves) or product emission/con-
sumption (transient HRR and concentration curves). Indeed, when
external heat flux increases from 11 to 60 kW m  2, the SMLR 6. Conclusion
increases from 11 to 32 g s  1 m  2 and the HRR increases from
275 to 780 kW m  2, respectively. Moreover, external heat flux This study has dealt with the experimental characterization of
has an influence on thermal degradation and product evolution thermal degradation and the combustion of a non-fire-retarded
durations, which are faster for higher external heat (  10 min at black PMMA. Experiments were carried out using a CC device
60 kW m  2 vs.  45 min at 11 kW m  2). under piloted ignition in order to characterize the external
As seen previously (Sections 3.3), in our experimental conditions heat flux influence on the degradation process and on gaseous
in the CC, the PMMA sample thermal decomposition seems to take emissions. The theoretical thermal properties of PMMA degrada-
place according to four time ranges – more and less significant and tion have been calculated. In addition, the exhaust gas analysis
separated, depending on incident heat flux – which are deduced from was performed using an FTIR spectrometer and a HORIBA gas
Fig. 3 and summarized in Table 3. From transient evolutions of the analyzer for heat fluxes ranging from 10 to 60 kW/m2 (with an
main released product concentrations, MLRs and HRRs, the four main experimental CHF equal to 11 kW m  2). The gaseous species
time ranges where thermal decomposition of black PMMA takes with the main concentrations and main emission yields were

Table 7
Atomic balance between the solid (source) and the exhaust gas (emissions) measured for the 11 irradiance levels studied.

External PMMA mass (g) Initial atomic mass (g) Atomic balance in exhaust gases (g)
heat flux (kW m  2)
C (from PMMA) H (from PMMA) C (from CO and CO2) H (from H2O)

11 168.7 72.5 99.7 7 1.5 13.3 7 0.2 94.1 718.8 13.5 7 4.5
12 171.6 101.4 13.6 93.1 718.6 14.5 7 4.8
15 171.6 72.5 101.4 7 1.5 13.6 7 0.2 100.1 720.0 14.3 7 4.7
20 169.2 73.6 100.07 2.1 13.4 7 0.3 100.7 720.1 13.1 7 4.3
25 173.3 72.5 102.4 7 1.5 13.7 7 0.2 98.3 719.7 12.6 7 4.2
30 167.5 72.6 99.07 1.5 13.2 7 0.2 99.1 719.8 11.8 7 3.9
35 168.0 72.1 99.3 7 1.2 13.3 7 0.2 100.4 720.1 12.0 7 3.9
40 168.2 71.1 99.4 7 0.7 13.3 7 0.1 101.6 720.3 10.9 7 3.6
45 170.4 72.6 100.77 1.5 13.5 7 0.2 103.4 720.7 12.3 7 4.1
50 171.6 73.0 101.4 7 1.8 13.6 7 0.2 104.2 720.8 11.6 7 3.8
55 168.8 72.5 99.8 7 1.5 13.3 7 0.2 101.8 720.4 11.5 7 3.8
60 168.7 72.5 99.7 7 1.5 13.3 7 0.2 102.6 720.5 11.0 7 3.6
Mean 169.6 72.9 100.47 1.5 13.4 7 0.2 99.9 720.0 12.4 7 4.1
460 J. Luche et al. / Fire Safety Journal 46 (2011) 451–461

CO2, CO, H2O, NO and O2, regardless of the irradiance level [2] D. Hopkins Jr, J.G. Quintiere, Material fire properties and predictions for
studied. The results have shown that emission yields presented thermoplastics, Fire Safety Journal 26 (1996) 241–268.
[3] B.T. Rhodes, Burning Rate and Flame Heat Flux for PMMA in the Cone
quasi-constant values (measurement and calculation uncertain- Calorimeter, NIST-GCR-95-664, National Institute of Standards and Technol-
ties taken into account) and were slightly influenced by an ogy, 1994.
external heat flux increase. [4] J.G. Quintiere, B.T. Rhodes, Fire Growth Models for Materials, NIST-GCR-94-
647, National Institute of Standards and Technology, 1994.
The main product evolutions have been correlated to the [5] T.-H. Tsai, M.-J. Li, I.-Y. Shih, R. Jih, S.-C. Wong, Experimental and numerical
evolutions of the HRR and the SMLR. From the HRR and SMLR, study of autoignition and pilot ignition of PMMA plates in a cone calorimeter,
EHC  24.5 kJ g  1 was computed for each irradiance level. Since Combustion and Flame 124 (2001) 466–480.
[6] S. Agrawal, A. Atreya, Wind-aided flame spread over an unsteady vaporizing
the experiments were carried out in an atmospheric area without
solid, in: 24th Symposium (International) on Combustion, The Combustion
external ventilation, the oxygen supplied to the flame was only Institute, Pittsburgh, PA, 1992, vol. 24, pp. 1685–1693.
piloted by convection. Then, oxidative reactions were mainly [7] C. Vovelle, J.-L. Delfau, M. Reuillon, J. Bransier, N. Laraqui, Experimental and
piloted by flame temperature. As expected, the results obtained numerical study of the thermal degradation of PMMA, Combustion Science
and Technology 53 (1987) 187–201.
confirm the influence of irradiance levels on the degradation of [8] J.L. Jackson, Direct Measurement of Heat of Gasification for Polymethyl-
the PMMA sample (i.e. on SMLR, HRR and product amounts in the methacrylate, National Institute of Standards and Technology, 1986 NISTIR
gaseous phase). 88-3809, September.
[9] P.A. Beaulieu, Flammability characteristics at heat flux levels up to 200 kW/
Using Thermogravimetric Analysis or a Tubular Furnace, m2 and the effect of oxygen on flame heat flux, Ph.D. Thesis. Worcester
where the reactor temperature was increased at a given heating Polytechnic Institute, 2005.
rate, some previous studies [23,24,50–56] have shown that the [10] R. Magee and R. Reitz, Extinguishment of radiation augmented fires by water
sprays, in: 15th Symposium (International) on Combustion, The Combustion
non-charring black PMMA polymer was thermally degraded Institute, Pittsburgh, PA, vol. 15, 1974, pp. 237–247.
according to the following four-step reaction mechanism (where [11] C. Vovelle, R. Akrich, J.L. Delfau, Mass loss rate measurements on solid
the term ‘‘gas’’ represents CO, CO2, H2O): materials under radiative heating, Combustion Science and Technology 36
(1984) 1–18.
[12] J.E.J. Staggs, The heat of gasification of polymers, Fire Safety Journal 39 (2004)
– Depolymerization of PMMA to form MMA-derived monomers 711–720.
[23,24,50–56] by pyrolysis or oxidation reactions, [13] A. Tewarson, R.F. Pion, Flammability of plastics—I. Burning intensity, Com-
bustion and Flame 26 (1976) 85–103.
Ox:=Py: [14] A. Tewarson, SFPE Handbook of Fire Protection Engineering, NFPA, 2002,
PMMA!MMA-derived compoundsþ Gas
Generation of Heat and Chemical Compounds in Fire, pp. 3–82 [Section 3,
Chapter 4].
[15] J.R. Ebdon, B.J. Hunt, P. Joseph, C.S. Konkel, D. Pric, K. Pyrah, T.R. Hull,
– Oxidation of the MMA-derived compounds and their trans- G.J. Milnes, S.B. Hill, C.I. Lindsay, J. McCluskey, I. Robinson, Thermal degrada-
formations into oxygenated compounds (alcohol, carboxylic tion and flame retardance in copolymers of methyl methacrylate with
acid, etc.) [23,24,47–53] inside the solid matrix, diethyl(methacryloyloxymethyl)-phosphonate, Polymer Degradation and
Stability 70 (2000) 425–436.
O2 [16] D. Price, K. Pyrah, T.R. Hull, G.J. Milnes, J.R. Ebdon, B.J. Hunt, P. Joseph, Flame
MMAderived compounds!Tar=Oxygenated compounds þ Gas
retardance of poly(methyl methacrylate) modified with phosphorus-contain-
ing compounds, Polymer Degradation and Stability 77 (2002) 227–233.
[17] A. Tewarson, F.H. Jiang, T. Morikawa, Ventilation-controlled combustion of
– Oxidation of the oxygenated compounds to form exhaust polymers, Combustion and Flame 95 (1993) 151–169.
gases and small gaseous compounds (formaldehyde, metha- [18] J.L. Consalvi, Y. Pizzo, B. Porterie, Numerical analysis of the heating process in
nol, acetone, acetylene, etc.), upward flame spread over thick PMMA slabs, Fire Safety Journal 43 (2008)
351–362.
O2 [19] R.E. Lyon, J.G. Quintiere, Criteria for piloted ignition of combustible solids,
Tar=Oxygenated compounds!Small gaseous compoundsþ Gas
Combustion and Flame 151 (2007) 551–559.
[20] J. Zhang, X. Wang, F. Zhang, A.R. Horrocks, Estimation of heat release rate for
polymer–filler composites by cone calorimetry, Polymer Testing 23 (2004)
– Small gaseous compound oxidation into gas, 225–230.
O2 [21] D.J. Irvine, J.A. McCluskey, I.M. Robinson, Fire hazards and some common
Small gaseous compounds!Gas polymers, Polymer Degradation and Stability 67 (2000) 383–396.
[22] V. Babrauskas, in: V. Babrauskas, S.J.. Grayson (Eds.), The Effects of FR Agents
on Polymer Performance, Plastics, Part B. Chapter 12: Heat Release in Fires,
Elsevier Applied Science, NY, 1992, pp. 423–446.
[23] W.R. Zeng, S.F. Li, W.K. Chow, Preliminary studies on burning behavior of
This mechanism probably takes place during the entire PMMA polymethylmethacrylate (PMMA), Journal of Fire Sciences 20 (2002)
thermal degradation process, and the four steps occur continu- 297–317.
ously on the whole duration of the CC experiments as soon as [24] W.R. Zeng, S.F. Li, W.K. Chow, Review on chemical reactions of burning
poly(methyl methacrylate) PMMA, Journal of Fire Sciences 20 (2002)
ignition starts. Using numerical simulation, the next step of this
401–433.
work will be to calculate the Arrhenius parameters of each [25] B.S. Kang, S.G. Kim, J.S. Kim, Thermal degradation of poly(methyl methacry-
reaction by using a parameter optimization method (e.g. genetic late) polymers: kinetics and recovery of monomers using a fluidized bed
algorithm [57–59]) to demonstrate the validity of this PMMA reactor, Journal of Analytical and Applied Pyrolysis 81 (2008) 7–13.
[26] Y.H. Hu, C.Y. Chen, Study of the thermal behaviour of poly(methyl metha-
thermal degradation mechanism in CC experiments under well- crylate) initiated by lactams and thiols, Polymer Degradation and Stability 80
ventilated air atmospheres. (2003) 1–10.
[27] Y.H. Hu, C.Y. Chen, The effect of end groups on the thermal degradation of
poly(methyl methacrylate), Polymer Degradation and Stability 82 (2003)
81–88.
Acknowledgments [28] ISO 5 660, Reaction to fire tests—heat release, smoke production and mass
loss rate—Part 1 Heat release rate (cone calorimeter method), 2002.
[29] ISO 17554:1998 International Standard – Fire Tests – Reaction to Fire – Mass
The authors are grateful for the financial support of the French loss measurement.
MAIF foundation (‘‘Fondation MAIF’’) and the COMPFEU program [30] T. Hakkarainen, E. Mikkola, J. Laperre, F. Gensous, P. Fardell, Y. Le Tallec,
to this work. C. Baiocchi, K. Paul, M. Simonson, C. Deleu, E. Metcalfe, Smoke gas analysis by
Fourier transform infrared spectroscopy—summary of the SAFIR project
results, Fire and Materials 24 (2000) 101–112.
References [31] ISO 19702, Toxicity testing of fire effluents—guidance for analysis of gases
and vapours in fire effluents using FTIR gas analysis, ISC 13.220.01, 2006.
[32] T. Hakkarainen, E. Mikkola, J. Laperre, F. Gensous, P. Fardell, Y. Le Tallec,
[1] B.T. Rhodes, J.G. Quintiere, Burning rate and flame heat flux for PMMA in a C. Baiocchi, K. Paul, M. Simonson, C. Deleu, E. Metcalfe, Smoke Gas Analysis
cone calorimeter, Fire Safety Journal 26 (1996) 221–240. by Fourier Transform Infrared Spectroscopy, The SAFIR Project, European
J. Luche et al. / Fire Safety Journal 46 (2011) 451–461 461

Commission Research Program Contract Number SMT 4-CT96-2136— [47] C. Hugett, Estimation of rate of heat release by means of oxygen consumption
Technical Research Centre of Finland, Copyright Valtron Teknillinen Tutki- measurements, Fire and Materials 12 (2) (1980) 61–65.
muskesus, 1999. /http://www.vtt.fi/rte/firetech/research/htmlS Smoke ana- [48] V. Babrauskas, Development of the Cone Calorimeter—A Bench-Scale Heat
lysis by Fourier transform infra-red spectroscopy (SAFIR). Release Rate Apparatus Based on Oxygen Consumption.NBSIR82-2611,
[33] P. Fardell, E. Guillaume, Sampling and measurement of toxic fire effluents, in: National Bureau of Standard, Gaithersburg, USA, 1982.
Fire Toxicity [Chapter 11], ISBN 1 84569 502 X–ISBN-13: 978 1 84569 502 6, [49] V. Babrauskas, Development of the cone calorimeter—a bench-scale heat
March 2010. release rate apparatus based on oxygen consumption, Fire and Materials 8
[34] L. Speitel, Fourier transform infrared analysis of combustion gases, Journal of (1984) 81–95.
Fire Sciences 20 (2002) 349–371.
[50] M. Comuce, T. Rogaume, F. Richard, T. Fateh, J. Luche, P. Rousseaux, Kinetics
[35] L. Speitel, DOT/FAA/AR-01/88 Fourier transform infrared analysis of combus-
and mechanisms of the thermal degradation of poly(methyl methacrylate by
tion gases, Final Report, October 2001.
thermogravimetry and Fourier-transform infrared spectroscopy analysis, in:
[36] A. Tewarson, D. Ogden, Fire behavior of polymethylmethacrylate, Combus-
tion and Flame 89 (1992) 237–259. Sixth International Seminar on Fire and Explosion Hazards, Weetwood Hall,
[37] M.A. Delichatsios, Piloted ignition times, critical heat fluxes and mass Leeds, UK, April 11–16, 2010.
loss rates at reduced oxygen atmospheres, Fire Safety Journal 40 (2005) [51] H. Arisawa, T.B. Brill, Kinetics and mechanisms of flash pyrolysis of poly
197–212. (methyl methacrylate) (PMMA), Combustion and Flame 109 (1997) 415–426.
[38] M.J. Spearpoint, J.G. Quintiere, Predicting the piloted ignition of wood in the [52] G. Lopez, M. Artetxe, M. Amutio, G. Elordi, R. Aguado, M. Olazar, J. Bilbao,
cone calorimeter using an integral model—effect of species, grain orientation Recycling poly-(methyl methacrylate) by pyrolysis in a conical spouted bed
and heat flux, Fire Safety Journal 36 (2001) 391–415. reactor, Chemical Engineering and Processing: Process Intensification 49 (10)
[39] Janssens, M.L.; Kimble, J.; Murphy, D., Computer tools to determine material (2010) 1089–1094.
properties for fire growth modeling from cone calorimeter data, in: Proceed- [53] J.E. Brown, T. Kashiwagi, Gas phase oxygen effect on chain scission and
ings of Fire and Materials 2003, 8th International Conference, Conference monomer content in bulk poly(methy1methacrylate) degraded by external
Papers, Organised by Interscience Communications Limited. January 27–28, thermal radiation, Polymer Degradation and Stability 52 (1996) 1–10.
2003, San Francisco, CA, pp. 377–387. [54] Z. Jiang, W.K. Chow, J. Tang, S.F. Li, Preliminary study on the suppression
[40] V. Babrauskas, R. Peacock, Heat release rate: the single most important chemistry of water mists on poly(methyl methacrylate) flames, Polymer
variable in fire hazard, Fire Safety Journal 18 (1992) 255–272. Degradation and Stability 86 (2004) 293–300.
[41] V. Babrauskas, Ignition Handbook, Fire Sciences Publishers, 0-9728111-3-3, [55] W. Kaminsky, M. Predel, A. Sadiki, Feedstock recycling of polymers by
2003.
pyrolysis in a fluidised bed, Polymer Degradation and Stability 85 (2004)
[42] V. Babrauskas, S. Grayson, Heat Release in Fires, Elsevier Science Publishers
1045–1050.
Ltd, Essex, England, 1992.
[56] K. Smolders, J. Baeyens, Thermal degradation of PMMA in fluidised beds,
[43] M.L. Janssens, Measuring Rate of Heat Release by Oxygen Consumption, Fire
technology 27 (1991) 234–249. Waste Management 24 (2004) 849–857.
[44] P.A. Enright, Heat release and the combustion behaviour of upholstered [57] C. Lautenberger, G. Rein, C. Fernandez-Pello, The application of a genetic
furniture, Ph.D. Thesis, University of Canterbury, 1999. algorithm to estimate material properties for fire modeling from bench-scale
[45] M. Janssens, SFPE Handbook of Fire Protection Engineering, NFPA, 2002 fire test data, Fire Safety Journal 41 (2006) 204–214.
[Section 3, Chapter 4, Calorimetry], pp. 38–62. [58] C. Lautenberger, C. Fernandez-Pello, A model for the oxidative pyrolysis of
[46] W. Thornton, The relation of oxygen to the heat of combustion of organic wood, Combustion and Flame 156 (2009) 1503–1513.
compounds, Philosophical Magazine and Journal of Science 33 (1917) [59] C. Lautenberger, C. Fernandez-Pello, Generalized pyrolysis model for com-
196–203. bustible solids, Fire Safety Journal 44 (2009) 819–839.

You might also like